• No results found

Under the assumption that this assertion is true (Conjecture 3.3), we say that (Sn) satisfies the Restricted Prime Number Theorem

N/A
N/A
Protected

Academic year: 2021

Share "Under the assumption that this assertion is true (Conjecture 3.3), we say that (Sn) satisfies the Restricted Prime Number Theorem"

Copied!
32
0
0

Loading.... (view fulltext now)

Full text

(1)

AMONG THE SUMS OF THE FIRST 2n PRIMES

ROMEO ME ˇSTROVI ´C

ABSTRACT. Let pnbe nth prime, and let (Sn)n=1:= (Sn) be the sequence of the sums of the first 2n consecutive primes, that is, Sn=P2n

k=1pkwith n = 1, 2, . . .. Heuristic arguments supported by the corresponding computational results suggest that the primes are distributed among sequence (Sn) in the same way that they are distributed among positive integers. In other words, taking into account the Prime Number Theorem, this assertion is equivalent to

#{p : p is a prime and p = Sk for some k with 1 ≤ k ≤ n}

∼#{p : p is a prime and p = k for some k with 1 ≤ k ≤ n} ∼ log n

n as n → ∞, where |S| denotes the cardinality of a set S. Under the assumption that this assertion is true (Conjecture 3.3), we say that (Sn) satisfies the Restricted Prime Number Theorem.

Motivated by this, in Sections 1 and 2 we give some definitions, results and examples concerning the generalization of the prime counting function π(x) to increasing positive integer sequences.

The remainder of the paper (Sections 3-7) is devoted to the study of mentioned se- quence (Sn). Namely, we propose several conjectures and we prove their consequences concerning the distribution of primes in the sequence (Sn). These conjectures are mainly motivated by the Prime Number Theorem, some heuristic arguments and related computational results. Several consequences of these conjectures are also established.

1. INTRODUCTION, MOTIVATION ANDPRELIMINARIES

An extremely difficult problem in number theory is the distribution of the primes among the natural numbers. This problem involves the study of the asymptotic behavior of the counting function π(x) which is one of the more intriguing functions in number theory. The function π(x) is defined as the number of primes ≤ x. For elementary methods in the study of the distribution of prime numbers, see [12].

Although questions in number theory were not always mathematically en vogue, by the middle of the nineteenth century the problem of counting primes had attracted the at- tention of well-respected mathematicians such as Legendre, Tch´ebychev, and the prodi- gious Gauss.

A query about the frequency with which primes occur elicited the following response:

I pondered this problem as a boy, and determined that, at aroundx, the primes occur with density1/ log x–C. F. Gauss (letter to Encke, 24 December 1849). Gauss wrote:

This remark of Gauss can be interpreted as predicting that

#{primes ≤ x} ≈

bxc

X

n=2

1 log n ≈

Z x 2

dt

log t = Li(x).

2010 Mathematics Subject Classification. Primary 11A41, Secondary 11A51, 11A25.

Keywords and phrases: distribution of primes, prime counting function, Prime Number Theorem, prime-like sequence, ω-Restricted Prime Number Theorem.

1

(2)

Studying tables of primes, C. F. Gauss in the late 1700s and A.-M. Legendre in the early 1800s conjectured the celebrated Prime Number Theorem:

π(x) = |{p ≤ x : p prime}| ∼ x log x (here, as always in the sequel, |S| denotes the cardinality of a set S).

This theorem was proved much later ([8, p. 10, Theorem 1.1.4]; for its simple analytic proof see [31] and [46], and for its history see [3], [21], [22] and [28, p. 21]. Briefly, π(x) ∼ x/ log x as x → ∞, or in other words, the density of primes p ≤ x is 1/ log x;

that is, the ratio π(x) : (x/ log x) converges to 1 as x grows without bound. Using L’Hˆopital’s rule, Gauss showed that the logarithmic integral Rx

2 dt/ log t, denoted by Li(x), is asymptotically equivalent to x/ log x. Recall that Gauss felt that Li(x) gave better approximations to π(x) than x/ log x for large values of x.

Though unable to prove the Prime Number Theorem, several significant contribu- tions to the proof of Prime Number Theorem were given by P. L. Chebyshev in his two important 1851–1852 papers ([6] and [7]). Chebyshev proved that there exist positive constants c1 and c2 and a real number x0 such that c1x/ log x ≤ π(x) ≤ c1x/ log x for x > x0. In other words, π(x) increases as x log x. Using methods of complex analysis and the ingenious ideas of Riemann (forty years prior), this theorem was first proved in 1896, independently by J. Hadamard and C. de la Vall´ee-Poussin (see e.g., [33, Section 4.1]).

A generalized prime system (or g-prime system) G is a sequence of positive real num- bers q1, q2, q3, . . . satisfying 1 < q1 ≤ q2 ≤ · · · ≤ qn ≤ qn+1 ≤ · · · and qn → ∞ as n → ∞. From these can be formed the system N of generalized integers or Beurling integers; that is, the numbers of the form qk11qk2l· · · qkll, where l ∈ Nand k1, k2, . . . , kl

N0 :=NS{0}. Notice that N denotes the multiplicative semigroup generated by G, and it consists of the unit 1 together with all finite power-products of g-primes, arranged in increasing order and counted with multiplicity.

Clearly, this system generalizes the notion of primes and positive integers obtained from them. Such systems were first introduced by A. Beurling [5] and have been studied by many authors since then (see in particular [4], [2], [11], [13], [25], [32] and [47]). In particular, Nyman [32] and Malliavin [25] sharpened Beurling’s results in various ways.

Much of the theory concerns connecting the asymptotic behaviour of g-prime counting functionand g- counting function πG(x) and NG(x), defined on [1, ∞) respectively by

πG(x) = X

q∈G,q≤x

1 and NG(x) = X

n∈N ,n≤x

1,

where in the first sum the summation is taken over all g-primes, counting multiplicities.

Similarly, for the second sumP

n∈N ,n≤x1. Accordingly, we have

πG(x) = #{i : qi ∈ G and gi ≤ x} and NG(x) = #{i : ni ∈ N and ni ≤ x}.

If G = {a1, a2, . . . , an, an+1, . . .} = (an)n=1 is a sequence such that a1 ≤ a2 ≤ · · · ≤ an ≤ an+1 ≤ · · · , then obviously, we have πG(an) = n for each n ∈N.

In 1937 Beurling proved [5, Th´eor`eme IV] that if NG satisfies the asymptotic relation NG(x) = Ax + O(x/ logγx) with some constants A > 0 and γ > 3/2, then the number of qn’s such that qn≤ x is equal to x/ log x + o(x/ log x), i.e.,

πG(x) = x log x+ o

 x

log x

 .

(3)

In other words, the conclusion of the Prime Number Theorem (in the sequel, shortly writ- ten as PNT) is valid for such a system G (from this reason often called a Beurling prime number system). Beurling also gave an example in which NG(x) = Ax + O(x/ log3/2x) but PNT is not valid. This result was refined in 1969 by H. G. Diamond [9, Theorem (B)]. In 1970 Diamond [10] also proved that Beurling’s condition is sharp, namely, the PNT does not necessarily hold if γ = 3/2.

In particular, if G is a set P := {p1, p2, p3, . . .} of all primes 2 = p1 < p2 < p3 < · · · with the associated multiplicative semigroup N = N = {1, 2, 3, . . .}, then PNT states that

π(x) ∼ x

log x, as x → ∞, where π(x) is the usual prime counting function, that is,

π(x) = X

p prime, p<x

1.

As observed in [4, Introduction], the additive structure of the positive integers is not particularly relevant to the distribution of primes. Therefore, for a given g-prime sys- tem G defined above, it can be of interest to consider the distribution of g-primes (the elements in G) with respect to certain associated system of generalized integers without any algebraic (multiplicative) structure. This means that the associated system N to G defined above may be some subset of [1, +∞) which is not a multiplicative semigroup (generated by G).

In particular, here we mainly consider the case when G is an infinite set of primes and the associated system N to G is an increasing integer sequence (an)n=1. We focus our attention when G is a set of all primes whose associated system N is the sequence (an)n=1:= (P2n

i=1pi)n=1where 2 = p1 < p2 < · · · < pn< · · · are all the primes.

Let (G, N := (ak)k=1) be a pair defined above. Then we define its counting function NG,(ak)(x) as

NG,(ak)(x) = #{i : i ∈Nand ai ≤ x}.

Furthermore, the prime counting function for (G, N := (ak)k=1) is the function x 7→

πG,(ak)(x) defined on [1, ∞) as

(1) πG,(ak)(x) = #{q : q ∈ G and q = ai for some i with ai ≤ x}.

Heuristic and computational results show that for many “natural pairs” (G, N := (ak)k=1) the associated counting function NG,(ak)(x) has certain asymptotic growth as x → ∞.

Notice that for each n ∈Nwe have

(2) πG,(ak)(an) = #{q : q ∈ G and q = ai for some i with 1 ≤ i ≤ n}.

The normalizable prime counting function for (G, N = (ak)k=1) is the function (n, x) 7→

pG,(ak)(n, x) defined for (n, x) ∈N× [1, +∞) as (3) pG,(ak)(n, x) = log an

an πG,(ak)(x).

The above expression induces the companion sequence (bn)n=1of (ak)k=1defined as (4) bn= pG,(ak)(an) = log an

an πG,(ak)(an), n = 1, 2, . . . .

(4)

We also define another “companion” sequence (cn)n=1of (ak)k=1 defined as

(5) cn = an

(log n)(log an) = πG,(ak)(an)

bnlog n , n = 1, 2, . . . .

Here as always in the sequel, we will suppose that G is a set of all primes whose associated system N is a sequence (ak)k=1. For brevity, in the sequel the functions defined by (1), (2), (3) and (4) will be denoted by p(ak)(x), π(ak)(an), p(ak)(n, x) and p(ak)(an), respectively.

Definition 1.1. Let Ω be a set of all nonnegative continuous real functions defined on (1, +∞) and let (ak)k=1 := (ak) be an increasing sequence of positive integers. We say that (ak) is a prime-like sequence if there exists the function ω(ak) = ω ∈ Ω such that the function n 7→ π(ak)(an) defined by (2) is asymptotically equivalent to ω(n) as n → ∞.

Then we say that a sequence (ak) satisfies the ω-Restricted Prime Number Theorem.

In particular, if ω(x) ∼ x/ log x as x → ∞, then we say that a sequence (ak) satisfies the Restricted Prime Number Theorem (RPNT).

Proposition 1.2. Let (ak)k=1 be a positive integer sequence, and let(bn)n=1be its com- panion sequence defined by(4). Then

(6) lim sup

n→∞

bn≤ 1.

Proof. Taking the obvious inequality π(ak)(an) ≤ π(an) with n = 1, 2, . . . into (4) we get

bn≤ π(an) log an

an , n = 1, 2, . . . , which by the Prime Number Theorem immediately yields

lim sup

n→∞

bn ≤ lim

n→∞

π(an) log an

an = 1,

as desired. 

Proposition 1.3. Let (an) be a prime-like sequence with the associated function ω(x).

Then

(7) lim sup

n→∞

ω(n) π(an) ≤ 1.

This means thatω(n) grows slowly than π(an) as n → ∞.

Proof. Notice that the inequality (7) is equivalent to

(8) lim sup

n→∞

log an

an π(ak)(an) ≤ 1.

Since by the assumption, ω(n) ∼ π(ak)(an), the inequality (8) yields lim sup

n→∞

log an

an ω(n) ≤ 1,

whence, in view of the fact that log an/an∼ 1/π(an), immediately follows (7).  Remark 1.4. The inequality (6) is sharp since by the Prime Number Theorem (see Ex- ample 2.1), equality in (6) holds for the sequences ak = k with k = 1, 2, . . ..

(5)

Remark1.5. If a sequence (ak) satisfies the ω-Restricted Prime Number Theorem, then by (4) we have

(9) ω(n) ∼ π(ak)(an) = anbn

log an as n → ∞.

Remark1.6. Let (ak) be a sequence satisfying the ω- Restricted Prime Number Theorem.

Then clearly, ω(ak)(n) ≤ n for all sufficiently large n. Moreover, ω(ak)(n) ∼ n as n → ∞ if and only if the density of primes in a sequence (ak) is equal to 1.

Notice also that by the Prime Number Theorem, ωN(x) = x/ log x for the sequence of all positive integers N = {1, 2, . . . n, . . .}, that is, N satisfies the Restricted Prime Number Theorem (cf. Conjecture 3.3).

Here, as always in the sequel, P = (pn) := {p1, p2, . . . pn, . . .} will denote the set of all primes, where 2 = p1 < 3 = p2 < p3 < · · · < pn< · · · . Moreover, (an) will always denotes an infinite strictly increasing sequence of positive integers. Hence, for such a sequence must be an ≥ n for each n ∈N.

The remainder of the paper is organized as follows. In Section 2 we present five examples concerning the determination of the function ω(ak)(x) and a sequence (bn) associated to a given sequence (ak). In particular, we consider the sequence (ak)k=1 with ak= a + (k − 1)d, where a ≥ 1 and d > 1 are relatively prime integers.

In Section 3 we consider the distribution of primes in the sequence (Sn)n=1 whose terms are given by Sn = P2n

i=1pi, where pi is the ith prime. Heuristic arguments sup- ported by related computational results suggest the curious conjecture that the sequence (Sn) satisfies the Restricted Prime Number Theorem (Conjecture 3.3). In other words, this means that the primes are distributed amongst all the terms of the sequence (Sn) in the same way that they are distributed amongst all the positive integers. Under this con- jecture, we prove that if qkis the kth prime in (Sn)n=1, then qk ∼ 2k2log3k ∼ 2p2klog k as k → ∞ (Corollaries 3.6 and 3.7).

Assuming that Conjecture 3.3 is true, in Section 4 we give the asymptotic expression for the kth prime in the sequence (Sn) (Corollary 4.2); namely, qk ∼ 2k2log3k as k → ∞. This result is refined by Theorem 4.4. We also conjecture that bk log kc+1 ≤ m for each pair (k, m) of positive integers with k ≥ 1 and qk= Sm(Conjecture 4.6). Some consequences of Conjectures 3.3 and 4.6 are also presented.

Section 5 is devoted to the estimations of the values Mk (k = 1, 2, . . .) involving in the expression for qk from Theorem 4.4. We also propose some other conjectures concerning the sequences (Sk) and (Mk). Related consequences are also established.

The conjectures presented in this paper, as well as some their consequences, are mainly supported by some computational results given in Section 6. In particular, the number πn := k of primes in the set Sn := {S1, S2, . . . , Sn} for 38 values of n up to 109+ 5 · 108 are presented in Table 1. For such values k and the associated indices m such that qk= Sm, the corresponding approximate values of qk, Mk(together with lower and upper bounds of Mk), (k log k)/m and Sm

√k log k/(2m5/2log m) are also given in this table. Under the previous notations, related numerical results for qk/(2k2log3k), qk/(2m2log m) and two estimates involving qk which are discussed in Section 4, are given in Table 3. Some additional computational results, the conjectures and their con- sequences are also given in Section 6.

(6)

In the last Section 7 we propose the stroner (asymptotic) version of Conjecture 3.3 which coincides with well known form of Prime Number Theorem involving the function li(x).

Notice that similar considerations to those for the sequence (Sn) concerning alternat- ing sums of consecutive primes are given in [29].

2. EXAMPLES

Example 2.1. For the sequence (ak) with ak = k (k = 1, 2, . . .), we clearly have π(k)(n) = π(n), and hence

(10) bn= π(n) log n

n , n = 1, 2, . . . . By the Prime Number Theorem, from (10) we find that

(11) lim

n→∞bn = lim

n→∞

π(n) log n

n = 1.

Example 2.2. Let (ak) be a sequence of all primes, that is, ak = pk with k ∈ N :=

{1, 2, . . .}, where pk is the kth prime. Since by (1), π(pk)(pn) = π(pn) = n, substituting this into (4) yields

(12) bn= n log pn

pn , n = 1, 2, . . . .

Now applying to (12) the well known fact that pn ∼ n log n as n → ∞ (see, e.g., [30]), we find that

(13) lim

n→∞bn= 1.

Notice also that the known inequality pn > n log n with n ≥ 1 (see, e.g., [38, (3.10) in Theorem 3]) implies that bn< 1 for all n ≥ 1.

Example 2.3. Suppose that a and d are relatively prime positive integers. Then concern- ing Dirichlet’s theorem de la Vall´ee Poussin established (see, e.g., [35, p. 205]) that the number of primes p < x with p ≡ a(mod d) is approximately

(14) π(x)

ϕ(d) ∼ 1

ϕ(d) · x log x.

Here ϕ(n) is the Euler totient function defined as the number of positive integers not exceeding n and relatively prime to n. Note that the right hand side of (14) is the same for any a such that gcd(a, d) = 1. This shows that primes are in a certain sense uniformly distributed in reduced residue classes with respect to a fixed modulus. Notice that for a sequence (ak)k=1given by ak= a + (k − 1)d, (14) can be written as

(15) π(ak)(an) ∼ π(an)

ϕ(d) as n → ∞.

Inserting (15) together with π(ak) ∼ ak/ log akinto (4) immediately gives

(16) lim

n→∞bn= 1 ϕ(d) lim

n→∞

π(an) log an

an = 1

ϕ(d).

(7)

Then substituting (16) into (9), we obtain that for the associated function ωa,d(x) := ω(ak of the sequence (ak) there holds

(17)

ωa,d(n) ∼ π(ak)(an) anbn

ϕ(d) log an = a + (n − 1)d

ϕ(d) log(a + (n − 1)d) ∼ dn

ϕ(d) log n as n → ∞.

It follows that ωa,d(x) = dx/(ϕ(d) log x) for x ∈ (1, +∞).

Example 2.4. Let (an) be a sequence defined as an= 2pn−1, where pnis the nth prime.

The numbers anare called Mersene numbers. A prime that appears in the sequence (an) is called Mersenne prime. Namely, it is easy to show (see, e.g., [36, p. 28]) that if 2n− 1 is prime, then so is n. The greatest known Mersenne prime is 243112609 − 1 with the exponent 43112609 (12978169 digit number), and it is discovered in August 2008. This is in fact one between 45 known Mersenne primes, and so a45 ≤ 243112609− 1.

In 1980 H. Lenstra and C. Pomerance, working independently, came the conclusion that the probability that a Mersenne number 2p− 1 is prime is eγlog(ap)/(p log 2) with γ = 0.577216 . . . (the Euler-Mascheroni constant), where a = 2 if p ≡ 3(mod 4) and a = 6 if p ≡ 1(mod 4). Recall that the constant eγ = 1.781072 . . . is important in num- ber theory; namely, eγ = limn→∞log p1

n

Qn k=1

pk

pk−1 which restates the third of Mertens’

theorems ([27], also see [23, pp. 351–353, Theorem 428]). Then notice that the distri- bution of the log of the Mersenne primes is a Poisson Process (see [45]).

Accordingly to the above assumption given by Lenstra and Pomerance, if ak = 2qk−1, where (qk)k=1is a sequence of all primes ≡ 3( mod 4) (q1 = 3, q2 = 7, q3 = 11, . . .), for the associated function ω(3,4)(x) to (ak) we have that “the expected number” of primes between the first n terms of the sequence (qk) is

(18) ∼ ω(3,4)(n) ∼

n

X

k=1

eγlog(2qk)

qklog 2 as n → ∞.

Since qk ∼ p2k ∼ 2k log k, substituting this into (18) and using the well known asymp- totic formulaPn

k=11/k ∼ γ + log n as n → ∞, we get

ω(3,4)(n) ∼ eγ 2 log 2

n

X

k=2

log(4k log k)

k log k = eγ 2 log 2

n

X

k=2

log k + log 4 + log log k k log k

∼ eγ 2 log 2

n

X

k=2

1 k +

n

X

k=2

log 4 k log k +

n

X

k=2

log log k k log k

!

∼ eγ 2 log 2



(γ + log n) + log 4 Z n

2

dx x log x+

Z n 2

log log x x log x dx



(the changes log x = s and log log x = t

= eγ 2 log 2



log n + Z log n

log 2

ds s +

Z log log n log log 2

t dt



∼ eγ 2 log 2



log n + log log n + (log log n)2 2



as n → ∞.

(19)

This shows that ω(3,4)(x) = eγ(log x + log log x + (log log x)2/2) /(2 log 2), and hence π(ak)(an) ∼ eγ/(2 log 2) (log n + log log n + (log log n)2/2). Substituting this in (4),

(8)

where (bn) is the companion’ sequence of (ak), and using the fact that qn ∼ 2n log n, we find that

(20) bn ∼ eγn(log n)2

4n log n as n → ∞.

Similarly, under the above assumptions attributed by Lenstra and Pomerance, if a0k = 2rk − 1, where (rk)k=1 is a sequence of all primes ≡ 1(mod 4) (r1 = 5, r2 = 13, r3 = 17, . . .), then for the associated function ω(1,4)(x) to (a0k) and the companion sequence (b0n) of (a0k) the same relations (18)–(20) are satisfied.

Example 2.5. Let (ak)k=1be an increasing sequence of positive integers satisfying

(21) log ak

ak = o(k−1).

Then from (4) and the obvious fact that π(ak)(an) ≤ n for each n ∈N, we find that

(22) lim

n→∞bn= 0.

In particular, (22) holds for any sequence (ak) satisfying one of the following asymp- totics: an ∼ an with a fixed a > 1; an ∼ n logαn with α > 1; an ∼ nα with α > 1; or an ∼ nαlogβn with α ≥ 1 and β > 1.

Accordingly, we ask the following question.

Question 2.6. For what real numbers α ∈ (0, 1) there exists a sequence (ak) whose companion sequence(bn) defined by (4) satisfies the limit relation

lim sup

n→∞

bn = α?

3. DISTRIBUTION OF PRIMES IN THE SEQUENCE(Sn)WITHSn=P2n i=1pi Here, as always in the sequel, we consider the distribution of primes in the sequence (Sn)n=1 whose terms are given by Sn =P2n

i=1pi, where pi is the ith prime. Recall that the prime counting function π(x) is defined as the number of primes ≤ x.

Proposition 3.1. Let (Sn) be the sequence defined as Sn=P2n

i=1pi. Then asn → ∞,

(23) Sn ∼ 2n2log n

and

(24) π(Sn) ∼ n2.

Furthermore, ifx is a real number such that Sn≤ x < Sn+1, then

(25) n ∼

r x

log x as n → ∞.

Proof. Let (Sn0) be the sequence defined as Sn0 = Pn

i=1pi (this is Sloane’s sequence A007504 in [42]). By the Prime Number Theorem, we have (see, e.g., [43, page 5]),

Sn0 :=

n

X

i=1

pn

n

X

k=1

k log k ∼ Z n

1

x log x dx = x2 2 log x

n 1

− Z n

1

x2

2 (log x)0dx

∼n2log n

2 as n → ∞.

(26)

(9)

It follows from (26) that

(27) Sn = S2n0 ∼ 2n2log n,

which implies (23). By the Prime Number Theorem, from (27) we have (28) π(Sn) ∼ 2n2log n

log(2n2log n) ∼ 2n2log n

log 2 + 2 log n + log log n ∼ n2.

Finally, (25) immediately follows from (23). 

Remark3.2. For refinements of the estimate (23), see [15], [39] and [41, Theorem 2.3]).

We see from (23) that there are ∼ n2primes less than Sn. Using this fact, Z.-W. Sun [43, Remark 1.6] conjectured that the number of primes in the interval (Pn

i=1pi,Pn+1 i=1 pi) is asymptotically equivalent to n/2 as n → ∞. Under the validity of this conjecture, in particular it follows that the number of primes in the interval (Sn, Sn+1) is asymptotically equivalent to n as n → ∞. Moreover, we also believe that the “probability” thatP2n

i=1pi is a prime is 2n/p2n, which is ∼ 1/ log n because of p2n ∼ 2n log 2n. Notice that the

“probability” of a large integer n being a prime is also asymptotically equal to 1/ log n.

Furthermore, some computational results and heuristic arguments show that between these ∼ n2 primes which are less than Sn there are ∼ 2n/ log Sn ∼ n/ log n primes that belong to the set Sn := {S1, S2, . . . , Sn}. For example, if n = 108 then n/ log n = 108/ log 108 = 5428681.02, while from the second column of Table 1 of Section 6 we see that there are 5212720 primes in the set S108 (cf. Table 2 of Section 6). Accordingly, we propose the following curious conjecture which is basic in this paper.

Conjecture 3.3. The sequence (Sn) with Sn = P2n

i=1pi satisfies the Restricted Prime Number Theorem. In other words,

πn : = π(Sk)(Sn) = #{p : p is a prime and p = Si for some i with 1 ≤ i ≤ n}

∼ n

log n as n → ∞.

(29)

Let us recall that in all results of this section (Corollaries 3.4, 3.6, 3.7, 3.8 and 3.13) we assume the truth of Conjecture 3.3. In particular, Conjecture 3.3 implies Euclid’s theorem (on the infinitude of primes) for (Sn) as follows.

Corollary 3.4 (Euclid’s theorem for the sequence (Sn)). The sequence (Sn) contains infinitely many primes.

Remark 3.5. Notice that the sequence (Sn) is closely related to the Sloane’s sequence A013918 [42] containing all primes (in increasing order) equal to the sum of the first m primes for some m ∈ N(A013918 is in fact the intersection of A000040-the sequence of all primes and A007504-sum of first n primes). The first few terms of the sequence A013918 are: 2, 5, 17, 41, 197, 281, 7699, 8893, 22039; see the related link by T. D. Noe [42, A013918] which gives the table of the first 10000 terms of this sequence (10000th term is 402638678093). Notice also that the Sloane’s sequence A013916 in [42] asso- ciated to the sequence A013918 gives numbers n such that the sum of the first n primes is prime. The first few terms of this sequence are: 1, 2, 4, 6, 12, 14, 60, 64, 96 (see the related link by D. W. Wilson [42, A013918] which gives table of the first 10000 terms of this sequence (10000th term is 244906). Similarly, the second Sloane’s sequence A013917 ((an)) associated to A013918, is defined as: anis prime and sum of all primes

≤ anis prime. The first few terms of this sequence are: 2, 3, 7, 13, 37, 43, 281.

(10)

As a further application of Conjecture 3.3, here we obtain the asymptotic expression for the kth prime in the sequence (Sn).

Corollary 3.6 (The asymptotic expression for the kth prime in the sequence (Sn)). Let qk(k = 1, 2, . . .) be the kth prime in the sequence (Sn). Then

(30) qk ∼ 2k2log3k as k → ∞.

Proof. If for a pair (k, n) there holds qk = Sn, then by Conjecture 3.3, we have

(31) k ∼ n

log n as n → ∞,

so that n ∼ k log n, and hence log n ∼ log k as n → ∞. Inserting this into (23), we find that

qk = Sn ∼ 2n2log n ∼ 2(k log n)2log n = 2k2log3n ∼ 2k2log3k,

as desired. 

Corollary 3.7. Let qk(k = 1, 2, . . .) be the kth prime in the sequence (Sn). Then

(32) qk∼ 2p2klog k as k → ∞

and

(33) qk∼ pk2log2k as k → ∞.

Proof. From (30) and the fact that pk∼ k log k we find that qk∼ 2(k log k)2log k ∼ 2p2klog k, which proves (32).

Similarly, from (30) and pk2 ∼ k2log k2 = 2k2log k we find that qk∼ (k2log k2) log2k ∼ pk2log2k,

which implies (33). 

Furthermore, we have the following result.

Corollary 3.8. Let qkbe thekth prime in the sequence (Sn) with qk= Sn. Then

(34) lim

k→∞

k log k n = 1.

Proof. The asymptotic relation (31) implies that log n/ log k ∼ 1, which substituting in

(31) immediately gives (34). 

Motivated by some heuristic arguments and computations for some small integer val- ues d, we propose the following generalization of Conjecture 3.3.

Conjecture 3.9. For any fixed nonnegative integer d the sequence (Sn(d))n=1defined as Sn(d) = 2d + Sn = 2d +

2n

X

i=1

pi, n = 1, 2, . . .

satisfies the Restricted Prime Number Theorem. In other words, asn → ∞, πn(d) :=π(2d+Sk)(2d + Sn) = #{p : p is a prime and p = 2d + Si

for some i with 1 ≤ i ≤ n} ∼ n log n. (35)

(11)

For d = 0 this conjecture is in fact Conjecture 3.3 (cf. Sloane’s sequence A013918 mentioned above).

Remark 3.10. Conjecture 3.3 and the fact that by (23) Sn ∼ 2n2log n imply that the average difference between consecutive primes in the sequence (Sn) near to 2n2 is ap- proximately log(2n2) ∼ 2 log n.

Remark3.11. Numerous computational results concerning the sums of the first n primes (partial sums of consecutive primes) given by the Sloane’s sequence A007504 (here denoted as Sn0), and certain their curious arithmetical properties are presented in the following Sloane’s sequences in OEIS [42]: A051838, A116536, A067110, A067111, A045345, A114216, A024011, A077023, A033997, A071089, A083186, A166448, A196527, A065595, A165906, A061568, A066039, A077022, A110997, A112997, A156778, A167214, A038346, A038347, A054972, A072476, A076570, A076873, A077354, A110996, A123119, A189072, A196528, A022094, A024447, A121756, A143121, A117842, A118219, A131740, A143215, A161436, A161490, A013918 etc.

Since the sequence (Sn) is a subsequence of the sequence (Sn0) with Sn0 = Pn k=1pk

whose all terms with odd indices n are even integers, it follows that in accordance to Definition 1.1, Conjecture 3.3 is equivalent to

ω(S0

k)(n) ∼ n 2 log n.

Therefore, Conjecture 3.3 is equivalent with the following one.

Conjecture 3.3’. Let (Sn0) be a sequence defined as Sn0 =Pn

k=1pk,n = 1, 2, . . .. Then

(36) ω(S0

k)(x) = x

2 log x f or x ∈ (1, ∞).

Proposition 3.12. For each n ≥ 3 we have

(37) 1 ≤ Sn

2n2log n < 1 + log 2

log n +log log(2n) log n .

Proof. By Mandl’s inequality (see, e.g., [39], [15]), for each n ≥ 9 there holds

(38) Sn0 < n

2pn

(for a refinement of (38), see [17, the inequality 2.4]). Mandl’s inequality (38) with 2n instead of n becomes Sn < np2n with n ≥ 5. This inequality together with the known inequality (see, e.g., [38, p. 69])

p2n < 2n(log n + log 2 + log log(2n)) for all n ≥ 3 immediately yields

(39) Sn< 2n2(log n + log 2 + log log(2n)) for all n ≥ 5.

On the other hand, a lower bound for Sn0 can be obtained by using Robin’s inequality (see, e.g., [15, p. 51]) which asserts that for every n ≥ 2

(40) np[n/2] ≤ Sn0.

The inequality (40) with 2n instead of n and the inequality n log n ≤ pn with n ≥ 3 (see, e.g., [38, p. 69]) yield

(41) 2n2log n ≤ Sn for n ≥ 3.

(12)

The inequalities (39) and (41) immediately yield log n ≤ Sn

2n2 < log n + log 2 + log log(2n) for all n ≥ 5, or equivalently,

(42) 1 ≤ Sn

2n2log n < 1 + log 2

log n +log log(2n)

log n for all n ≥ 5.

The inequalities given by (42) coincide with these of (37) for n ≥ 5. A direct calculation chows that (37) is also satisified for n = 3 and n = 4. This completes the proof.  Corollary 3.13. Let qk = Sm be thekth prime in the sequence (Sn)n=1. Then for all k ≥ 3 there holds

(43) 2m2log m < qk < 2m2(log m + log(log(2m) + log 2).

Proof. The above inequalities coincide with (37) of Proposition 3.12 with n = m and

qk = Sm. 

Remark3.14. Z.-W. Sun [44, the case α = 1 in Lemma 3.1] showed that for all n ≥ 2 Sn0 > 2 +n2log n

2



1 − 1

2 log n

 , which with 2n instead of n becomes

Sn > 2 + 2n2



log n + log 2

√e



≈ 2 + 2n2(log n + 0.193147), whence it follows that

Sn

2n2log n > 1 + 0.193147

n .

The above inequality is stronger that the left hand side of the inequality (37). Accord- ingly, if qk = Sm, then the first inequality of (43) can be refined in the form

qk> 2 + 2m2(log m + 0.193147) for all k ≥ 3.

On the other hand, combining the inequalities (46) and (47) from the next section with the inequalities Sn > 2npn and Sn < np2n (given in proof of Proposition 3.12), respectively, we immediately obtain the following refinement of Proposition 3.12.

Proposition 3.15. For each n ≥ 3 there holds Sn

2n2log n ≥ 1 + log log n − 1

log n + log log n − 2.2 log2n , and for eachn ≥ 344192, we have

Sn

2n2log n ≤ 1 + log log(2n) + log 2 − 1

log n + log log(2n) − 2 (log n) log(2n).

Remark 3.16. If qk = Sm, then in view of the first inequality of Proposition 3.15, the first inequality of (43) may be replaced by the following one:

qk> 2m2



log m + log log m − 1 +log log m − 2.2 log m



for all k ≥ 3.

(13)

Remark3.17. The inequalities (38) and (40) and the asymptotic expression pn∼ n log n show that the average of the first n primes is asymptotically equal to (n log n)/2 (cf.

Sloane’s sequence A060620 in [42]), that is, Sn0

n ∼ n log n

2 as n → ∞.

Conjecture 3.3 suggests the fact that for the sequence (Sn) would be valid the ana- logues of some other classical results and conjectures closely related to the Prime Num- ber Theorem and Riemann Hypothesis. In particular, if Q = {q1, q2, . . . , qk, . . .} is a set of all primes q1 < q2 < . . . < qk < · · · in the sequence (Sn), it can be of interest to establish the asymptotic expression for qkas k → ∞.

Finally, heuristic arguments, some computational results and Conjecture 3.3 lead to the follwing its generalization (cf. Sloane’s sequence A143121 - triangle read by rows, T (n, k) =Pn

j=kpj, 1 ≤ k ≤ n; see the columns in Example of this sequence).

Conjecture 3.18. For any fixed positive integer k, let (Sn(k)) := (Sn(k))n=1 be the se- quence whosenth term is defined as

Sn(k) =

2n+1

X

i=1

pi+k, n ∈ N.

Then the sequence(Sn(k)) satisfies the Restricted Prime Number Theorem.

For example, there are 78498 (resp. 664579) primes less than 106 (resp. 107), while the computations show that among the first 106 (resp. 107) terms of the sequences (Sn), (Sn(k)) with k = 1, 2, . . . , 12 there are 69251 (resp. 594851), 69581 (resp. 594377), 68844 (resp. 593632), 68883 (resp. 595733), 69602 (resp. 596609), 69540 (resp.

596558), 69414 (resp. 595539), 69317 (resp. 594626), 69455 (resp. 595474), 69268 (resp. 594542), 68891 (resp. 593807), 69251 (resp. 594383), 69564 (resp. 595270) primes, respectively.

4. THE ASYMPTOTIC EXPRESSION FOR THEkTH PRIME IN THE SEQUENCE(Sn) As an easy consequence of the Prime Number Theorem, it can be deduced that pn ∼ n log n as n → ∞ (see, e.g., [26]). Furthermore, a particular asymptotic expansion for pn(see [26] or [34, the equality (66) of Section 6]; also see Sloane’s sequence A200265) yields

(44) pn = n



log n + log log n + O log log n log n



. It is also known that (see [16] and [38, p. 69])

(45) n(log n + log log n − 1) < pn< n(log n + log log n).

A more precise work about this can be found in [37] and [40] where related results are as follows:

(46) n



log n + log log n − 1 + log log n − 2.2 log n



≤ pn for n ≥ 3 and

(14)

(47) pn ≤ n



log n + log log n − 1 + log log n − 2 log n



for n ≥ 688383.

The inequalities (46) and (47) immediately yield the following result.

Corollary 4.1. For each n ≥ 688383 the interval (48)

 n



log n + log log n − 1 + log log n − 2.2 log n

 , n



log n + log log n − 1 +log log n − 2 log n



contains at least one prime. Furthermore, the lenghtlnof this interval is

(49) ln= 0.2n

log n ∼ 0.2pn log2n.

As an application of Conjecture 3.3, we obtain the following result.

Corollary 4.2. Let qk be the kth prime in the sequence (Sn) with Sn = P2n

k=1pk. If qk = Sm, then under Conjecture3.3 there holds

(50) qk ∼ 2k2log3k ∼ 2m2log m ∼ pbk2log2kc as k → ∞, and

(51) lim

k→∞

k log k m = 1.

Proof. The first asymptotic relation of (50) coincides with (30) of Corollary 3.6. Further, by (37) of Proposition 3.12, we have

(52) qk= Sm ∼ 2m2log m.

Moreover, we have

(53) p[k2log2k] ∼ k2(log2k) log(k2log2k) ∼ 2k2log3k as k → ∞.

The last two asymptotic expressions of (50) follow from (52) and (53).

It remains to prove (51). If we suppose that (51) is not satisfied, then there exists ε > 0 and an infinite subsequence (kj, mj)j=1 of the sequence (k, m)k=1such that kjlog kj ≥ (1 + ε)mj for all j ∈Nor kjlog kj ≤ (1 − ε)mj for all j ∈ N. In the first case, using (50) for all sufficiently large j, we find that

m2jlog mj ∼ k2j log3kj ≥ (1 + ε)2m2jlog kj,

whence we immediately get log mj ≥ (1 + ε)2log kj, or equivalently, mj ≥ kjt with t = (1 + ε)2 > 1. From the previous inequality and the fact that t > 2 we have

m2j log mj > m2j ≥ kj2t  k2jlog3kj,

which contradicts the fact that by (50) 2k2log3k ∼ 2m2log m. In a similar way as in the first case, in the second case we find that mj ≤ ksj for a constant s = (1 − ε)2 < 1.

Then choosing a sufficiently large j0 such that log mj < m1/s−1j for all j > j0, in view of the fact that s + 1 < 2 we get

m2jlog mj < m1+1/sj ≤ kjs+1 kj2log3kj.

A contradiction, and therefore, (51) is true. 

(15)

Remark4.3. From (50) and pk ∼ k log k we see that

(54) qk

2k log2k ∼ k log k ∼ pk as k → ∞.

The above asymptotic expression together with the assumption that Conjecture 3.3 is true suggests the fact that for the sequence qk/(k log2k) would be satisfied the asymptotic expansion similar to (44), i.e.,

(55) qk

2k log2k = k(log k + log log k + Qk),

with some sequence (Qk). Motivated by (55), we establish the following asymptotic expression for the kth prime in the sequence (Sn).

Theorem 4.4 (The asymptotic expression for the kth prime in the sequence (Sn)). Let qk be thekth prime in the sequence (Sn) (k = 2, 3, . . .). Then under Conjecture 3.3 there exists a sequence(Mk) of positive rael numbers such that limk→∞Mk = 1 and

(56) qk = 2Mk5k2log2k(log k + log log k + 2 log Mk).

For the proof of Theorem 4.4 we will need the following result.

Lemma 4.5. Let Sm = qkbe thekth prime in the sequence (Sn). Then under Conjecture (3, 3) we have

(57) qk ∼ 2m2

m log m

√k log k as k → ∞.

Proof. First notice that, under notations of Lemma 4.5, Conjecture 3.3 yields (cf. (51) of Corollary 4.2)

(58) k ∼ m

log m as k → ∞.

Using (58), we find that (59) 2m2

m log m

√k log k ∼ 2m2

m log m r

m

1 −log log mlog m 

∼ 2m2log m as k → ∞.

The asymptotic relation (59) and the fact that by (50) of Corollary 4.2, qk = Sm

2m2log m immediately yield (57). 

Proof of Theorem4.4. Let (Ck)k=2 be a sequence of positive real numbers such that m(k) = m = Ckk log k with k ≥ 2 and qk = Sm. Then by (51) of Corollary 4.2, we have Ck → 1 as k → ∞. Taking m = Ckk log k into (57) of Lemma 4.5, as k → ∞ we obtain that

qk∼ 2 q

Ck5k5log5k(log k + log log k + log Ck)

√k log k

= 2Ck2p

Ckk2log2k(log k + log log k + log Ck) =: f (k, Ck).

(60)

Let (δk) be a positive real sequence such that

(61) qk = δkf (k, Ck) for each k ≥ 2.

(16)

Then from (60) we see that δk → 1 as k → ∞. For a fixed k ≥ 2 consider the equation fk(x) = δkf (k, Ck) which can be written in the form

(62) x2

x(log k + log log k + log x) = δkCk2p

Ck(log k + log log k + log Ck).

Notice that for any fixed integer k ≥ 2, the real function fk(x) defined as fk(x) = 2x2

x(log k + log log k + log x), x > 0,

satisfies the limit relations limx→+∞fk(x) = +∞ and limx→+0fk(x) = 0. From this it can be easily shown that for each integer k ≥ 2 the equation (62) has a positive real solution xk. Using the facts that limx→+∞Ck = limx→+∞δk=1, it can be easily show that limk→∞xk = 1. Then taking xk = Mk2(k = 2, 3, . . .), then limk→∞Mk = 1 and by (62) we find that fk(Mk2) = δkf (k, Ck) = qk, whence it follows that

qk = 2Mk5k2log2k(log k + log log k + 2 log Mk).

This proves (56) and the proof is completed. 

Computational results (cf. the eighth column of Table 1 of Section 6) suggest the additional relationship between k’s and m’s as follows.

Conjecture 4.6. For each pair (k, m) with k ≥ 1 and qk = Smwe have

(63) bk log kc + 1 ≤ m,

or equivalently,

(64) qk≥ Sbk log kc+1.

Furthermore, for eachk ≥ 104,

(65) m ≤ b1.4k log kc,

or equivalently,

(66) qk≤ Sb1.4k log kc.

Corollary 4.7. If the inequality (63) of Conjecture 4.6 is true, then for each k ≥ 1 there holds

(67) qk > 2k2(log2k)(log k + log log k).

Proof. Combining the inequality (63) with the inequality on the left hand side of (37) of Proposition 3.12, we find that

qk = Sm ≥ Sbk log kc+1 ≥ 2(bk log kc + 1)2log(bk log kc + 1)

> 2k2(log2k) log(k log k) = 2k2(log2k)(log k + log log k), (68)

as desired. 

Corollary 4.8. If the inequality (63) of Conjecture 4.6 is true, then Mk > 1 for each k ≥ 1, where (Mk) is the sequence defined by (56) of Theorem 4.4.

Proof. The assertion follows immediately from the inequality (67) and the expression

(56) for qkgiven by Theorem 4.4. 

Finally, in view of the data of the last column in Table 1 of Section 6 and some con- siderations presented above, we propose the following conjecture which is stronger than Corollary 4.7.

(17)

Conjecture 4.9. For every k ≥ 252028 with qk = Smthere holds

(69) qk > 2m2

m log m

√k log k .

In view of the well known inequality pk > k(log k + log log k), the following conjec- ture is also stronger than Corollary 4.7.

Conjecture 4.10. There exists k0Nsuch that for everyk ≥ k0 there holds qk > 2kpklog2k,

wherepkis thekth prime.

Remark4.11. The last column of Table 1 presented in Section 6 shows that qk ≈ 2m2

m log m

√k log k

is a “good” approximation for the kth prime sum qk. Notice that this approximation can be written as

qk≈ 2m2log m ·

r m

k log k,

where the valuespm/(k log k) slowly tend to 1 as k grows. In particular, from the last row of Table 1 of Section 6 we see that for m = 109− 2 (i.e., for k = 46388006) we have pm/(k log k) ≈ 1.105079. Hence, in view of the above approximation, we believe that for all values m up to 109 there holds

qk > 2.2m2log m.

Notice that some values of the sequences (Q0k) such that

Q0k = qk− 2k2(log2k)(log k + log log k) 2k2log2k log log k

and the sequence (Q00k) such that

Q00k= qk− 2(pk)2log k 2k2log2k log log k

are given in Table 3 of Section 6. Table 3 also shows that for almost all values m up to 109(i.e., for k ≤ 46388006) there holds

kpk> 1.1m2log m.

Finally, Table 2 of Section 6 leads to the following conjecture whose both parts are obviously stronger than Conjecture 4.6.

Conjecture 4.12. Let π(x) be the prime counting function, and let πnbe the number of primes in the set{S1, S2, . . . , Sn}. Then

πn < π(n) for for eachn ≥ 104 and

πn< n log n for eachn ≥ 105.

(18)

5. ESTIMATIONS OF VALUES MkFROMTHEOREM4.4

The computational results related to the search of primes in the sequence (Sn) given in the following section (Table 1) and some heuristic arguments suggest the fact that the sequence (Sn/(2n2log n))n=2 plays an important role for estimating the values Mk

(k = 1, 2, . . .) in the expression (56) for the kth prime qkin the sequence (Sn).

Here we first consider the sequence (Sn/(2n2)).

Proposition 5.1. The sequence (vn) defined as

(70) vn = Sn

2n2, n ∈N, is increasing forn ≥ 2.

Proof. Since Sn+1 = Sn+ p2n+1 + p2n+2, an easy calculation shows that rn < rn+1 is equivalent with

Sn

2n2 < p2n+1+ p2n+2 2(2n + 1) , which can be written as

(71) p2n+1+ p2n+2

2 > Sn 1 n + 1

2n2

 .

By a refinement of Mandl’s inequality due to Hassani [17], for every n ≥ 10 we have

(72) n

2pn

n

X

i=1

pi > 0.01659n2. Replacing n by 2n into (72) it becomes

(73) p2n− Sn

n > 0.06636n2 for all n ≥ 5.

Further, by the inequality (37) of Proposition 3.12 we have (74) log(2n) + log log(2n) > Sn

2n2 for all n ≥ 5.

By using Mathematica 8, it is easy to to prove the inequality (75) 0.06636n2 > log(2n) + log log(2n) for all n ≥ 8.

Finally, combining the inequalities (73), (74), (75) and the obvious inequality (p2n+1+ p2n+2)/2 > p2n immediately gives (71) for all n ≥ 8. This together with a direct verification that vn< vn+1for 2 ≤ n ≤ 8 concludes the proof.  Remark 5.2. Notice that the sequence (vn) defined by (70) is a subsequence of the se- quence (v0n) defined as

vn0 = 2Sn0

n2 := 2Pn i=1pi

n2 , n ∈ N;

namely, vn = v02nfor all n = 1, 2, . . .. Similarly as in the proof of Proposition 5.1, it can be shown that the sequence (v0n) is increasing for n ≥ 4.

Contrary to Proposition 5.1, we propose the following conjecture.

References

Related documents

The Milnor Conjecture was posed by John Milnor in 1970 to give a description of the Milnor K-theory ring of a field F of characteristic di fferent from 2 by means of the graded

Consequently, if the twin prime conjecture is true, then the number of strong primes is infinite..

What are the driving factors leading companies to request sales tax outsourcing services:. • Complexity of returns at the local level of tax (County

Delta Cultura Cabo Verde (DCCV) gathers and generates the data shown in the following graphs in the online M&amp;E software Infocus5. DCCV is able to use this software since

Senator Young felt the language used would make it mandatory rather than optional and Senator Swobe said this was not the intent, however he would check with

Goldbach’s weak conjecture which has been proven says that any odd number is the sum of just three prime numbers.. I have inserted a = 3 as one of these three prime numbers as

MENTALLY DISORDERED OFFENDER DISCHARGE PLANNING AIDS SERVICE PROVIDERS • FOOD • TRANSPORTATION • COUNSELING • SPIRITUAL • SUPPORT GROUPS COMMUNITY AFFILIATES HOUSING

• Plant &amp; Network Energy Performance • Water Distribution Optimization &amp; Loss Mgt •Stormwater management and Urban Flooding • Irrigation Management.